You are on page 1of 14

Int. J. Electrochem. Sci.

, 6 (2011) 1454 - 1467

International Journal of
ELECTROCHEMICAL
SCIENCE
www.electrochemsci.org

Inhibition Effect of Natural Artemisia Oils Towards Tinplate


Corrosion in HCl solution: Chemical Characterization and
Electrochemical Study
L. Bammou 1, M. Mihit 2, R. Salghi 1, A. Bouyanzer 3, S.S. Al-Deyab 4, L. Bazzi 2, B. Hammouti 3,*
1
Equipe Génie de l’Environnement et de Biotechnologie, ENSA, Université Ibn Zohr, BP 1136, 80000
Agadir, Morocco.
2
Laboratoire Environnement & Matériaux, BP 8106, Faculté des sciences, Université Ibn Zohr, 80000
Agadir, Morocco.
3
LCAE-URAC18, Université Mohammed Premier B.P. 4808, Oujda, Morocco.
4
Petrochemical Research Chair, Chemistry Department, College of Science, King Saud University,
P.O. Box 2455, Riyadh 11451, Saudi Arabia
*
E-mail: hammoutib@gmail.com

Received: 24 October 2011 / Accepted: 20 January 2011 / Published: 1 May 2011

The natural Artemisia essential oil (AO) was tested as a green inhibitor for tinplate in de-
aerated hydrochloride solution (0.5M HCl), by means of potentiostatic polarization, in the
range of temperature from 298 to 338 K. Polarization data obtained reveal that AO performed
excellently as a corrosion inhibitor for tinplate corrosion in hydrochloride acid media and acted
as a mixed type inhibitor. The inhibition efficiency increases with an increase in AO
concentration and attains 81% at 0.5 g/l at 298 K. Also, we ensured that the temperature favors
the dissolution of tinplate and becomes increasingly important when the temperature increases.
The Langmuir adsorption isotherm was tested for their fit to the experimental data. We also
present results of an in situ SEM study on the surface morphology of tinplate electrodes in
inhibited and uninhibited electrolytes. The parameters ( E *a , ΔH *a , ΔS*a , and ΔG *a ) were estimated
and discussed. The fundamental thermodynamic functions were used to glean important
information about Artemisia oil inhibitory behavior.

Keywords: Corrosion, green inhibitor, artemisia oil, tinplate

1. INTRODUCTION

Tinplate is a thin sheet iron or steel coated with tin to prevent rusting, used especially to
make cans and pots. The behaviour of tinplate in aqueous environments depends on several
Int. J. Electrochem. Sci., Vol. 6, 2011 1455

parameters such as the surface properties of the material, the temperature, pH and the
composition of the aggressive solution [1-5]. Several studies revealed that the degree of
dissolution of tin and the appearance of pitting on the surface of tinplate are more important in
the presence of Cl- anions [6-10]. Moreover, protein and amino acid in the canned food
usually contain chloride and sulphur resulting in the food can wall would be attacked seriously
and that food in the can being polluted as well. Also, the circumstance in the food can is
usually aggressive, and it is required that tinplate cannot seriously corrode during a food can’s
storage life [11].
The use of non-toxic inhibitors called green or eco-friendly environmental inhibitors is
one of the solutions possible to prevent the corrosion of the material and to keep the quality of
the food conserved. Natural plants have been added in the form of extracts, oils or pure
compounds, and may play major roles to keep the environment more healthy, safe, and under
pollution control. Among the various natural products, such as the lawsonia extract [12], black
peper [13], cedre oil [14], jojoba oil [15], phyllanthus amarus extract [16], fenugreek extracts
[17], artemisia oil [18], pennyroyal oil from mentha oulegium [19], hibiscus sabdariffa extract
[20], azadiracta indica [21], garcinia kola extract [22], eucalyptus oil [23,24] and thymus oil
[25]. All of which have been reported to be good inhibitors for metals and alloys in acidic
solutions.
The aim of this paper is to evaluate the effect of the addition of natural Artemisia
essential oil (AO) on the corrosion of tinplate alloy in hydrochloride solution (0.5M HCl)
taking account the variation of both medium temperature and inhibitor concentration.

2. MATERIALS AND METHODS

2.1. Plant collection and essential oil extraction

Artemisia was collected in the region of Agadir, Morocco. It was taxonomically


identified at the National Scientific Institute of Rabat (Laboratory of Botany, Department of
Plant Biology). A voucher specimen sample was deposited in the herbarium of the Laboratory,
the National School of Applied Sciences, Ibn Zohr University. Agadir, Morocco. The aerial
parts of the plant were air-dried in the laboratory at room temperature. A sample of 100 g was
subjected to water distillation for 2 hours using a Clevenger-type apparatus recommended by
the French Pharmacopoeia [26]. The yield was determined as grams over the 100 g of powder
analyzed in percentage, and is shown in table 1. The oil extracted was analyzed using a
Hewlett-Packard 5972 MS, fitted with a HP 5890 Series II GC, and controlled by a G1034C
Chemstation. A sample of 1µl was injected under the following conditions: DB1 fused silica
capillary column (20 m x 0.20 mm, film thickness 0.2 µm); carrier gas helium (0.6 ml/min);
injector temperature 523K; column temperature 323-523K at 276K/min; and MS electronic
impact 70 eV. The identification of the compounds was achieved by comparing retention
times and mass spectra with those of the published standards [27,28].
Int. J. Electrochem. Sci., Vol. 6, 2011 1456

2.2. Preparation of the electrode

The employed material in this study is the tinplate. It has been selected based on its diverse
uses in packing food and beverage cans. For electrochemical tests, the c o m m e r c i a l t i n p l a t e
s a m p l e s a r e i n t h e f o r m o f a surface disc of 0.5 cm 2 used as a work electrode that is rinsed
with acetone then with distilled water before plunging the electrode in the solution.

Table 1.Chemical composition of Artemisia essential oils.

Plant species Family Yield (%)a Major constituents (%)b

Artemisia Astéracées 1.5 Camphor, (C10H16O) 46.0


α-thuyone, (C10H16O) 33.2
β-thuyone, (C10H16O) 09.0
1,8-cineole, (C10H18O) 06.4
Camphene, (C10H16) 08.5
a
Grams of extracted oil as percentage over the 100 g of powder analysed.
b
Grams over total extracted oil in percentage.

2.3. Electrochemical tests

The electrochemical study was carried out using a potentiostat PGZ100 piloted by
Voltamaster soft-ware. This potentiostat is connected to a cell with three electrode thermostats
with double wall (Tacussel Standard CEC/TH). A saturated calomel electrode (SCE) and platinum
electrode were used as reference and auxiliary electrodes, respectively. The working electrode is in
form of disc from tinplate of the surface 0.5cm 2. Potentiodynamic polarization curves were plotted
from -700 to -500 mV/SCE at a polarization scan rate of 0.5 mV/s. Before all experiments, the
potential was stabilized at free potential during 30 min.
The electrochemical impedance spectroscopy (EIS) measurements are carried out with the
electrochemical system (Tacussel), which included a digital potentiostat model Voltalab PGZ100
computer at E cor after immersion in solution without bubbling. After the determination of steady-
state current at a given potential, sine wave voltage (10mV) peak to peak, at frequencies between
100 kHz and 10 mHz are superimposed on the rest potential. Computer programs automatically
controlled the measurements performed at rest potentials after 0.5, 2, 4 and 24 hours of exposure.
The impedance diagrams are given in the Nyquist representation.
The solution 0.5M HCl was prepared by dilution of concentrated HCl with double distilled
water. The solution tests are freshly prepared before each experiment by adding the oil directly to
the corrosive solution. The solution test is there after de-aerated by bubbling nitrogen. Gas pebbling
is maintained prior and through the experiments. Experiments are repeated three times to ensure the
reproducibility.
Int. J. Electrochem. Sci., Vol. 6, 2011 1457

2.4. Surface morphology

The surfaces (0.9 cm X 0.8 cm X 0.1 cm) of tinplate immersed for 1 day in 0.5M HCl
solution with and without 0.5 g/l of AO were analyzed by scanning electron microscopy (SEM)
JOEL JSM-5500 type.

3. RESULTS AND DISCUSSION

3.1. Artemisia oil (AO) analysis

GC-mass spectrum analyzes permit the identification of the composition of Artemisia


essential oil. The main components are listed in table 1. AO mainly contains camphor, α-
thuyone, β-thuyone, camphene, 1,8-cineole (Figure 1).

Camphor α-thuyone β-thuyone 1,8-cineole Camphene

Figure 1. Molecular structure of the major components of Artemisia essential oil.

3.2. Polarization curves

The composition of the medium and its temperature are essential parameters affecting
the phenomenon of the corrosion. Electrochemical steady state measurements are taken at
various temperatures (298-338K) in the absence and presence of Artemisia essential oil at
different concentrations (0.05 to 6g/l). The anodic and cathodic polarization curves obtained are
shown in Fig. 2 and the corresponding electrochemical parameters deduced from these curves
such as corrosion potential (Ecorr), corrosion current density (Icorr), anodic (βa) and cathodic (βc) Tafel
slopes estimated by using Tafel ruler are all given in table 2.
It’s observed that the anodic and cathodic branches of polarization curves show the same
characteristics with and without the addition of the inhibitor at all temperatures studied. Extrapolation
of this linear corrosion potential leads to the current density of corrosion. In anodic domain, an anodic
active passive transition is observed in uninhibited and inhibited acid.
Analysis of the polarization curves indicates that the presence of Artemisia oil tested leads to
decreases in anodic current densities where the cathodic current densities are nearly unchanged. The
inhibited systems were slightly shifted toward anodic potentials, emphasizing that the studied
compounds act as an anodic type inhibitor. We remark also that cathodic current potential curves give
Int. J. Electrochem. Sci., Vol. 6, 2011 1458

rise to parallel Tafel lines indicating that the hydrogen evolution reaction is activation controlled and
the addition of the inhibitor studied does not modify the mechanism of this process.

0,01
0.5 M HCl
1E-3

I(A.cm )
-2
1E-4

1E-5 298K
308K
318K
1E-6 328K
338K
1E-7
-0,70 -0,65 -0,60 -0,55 -0,50 -0,45
E(V/SCE)

0.5 M HCl +0.5 g/l Artemisia oil


1E-3

1E-4
I(A.cm )
2

1E-5 298K
308K
318K
1E-6 328K
338K
1E-7
-0.70 -0.65 -0.60 -0.55 -0.50 -0.45

E(V/SCE)
Figure 2. Polarization curves of tinplate in 0.5M HCl in the absence and presence of 0.5g/l of
Artemisia essential oil at various temperatures of the corrosive medium.

In a reducing acid where hydrogen is evolved from a tin surface, the exchange current is
relatively small. It may be increased and the cathodic Tafel slope decreased by providing local noble-
metal cathodes. In other words, the dissolution of tin is possible but the rate of corrosion may be very
slow. The hydrogen overpotential of tin is high [29].
As can be seen from table 2 that the increasing of temperature from 298 to 338 K acts on
all electrochemical parameter values; we observe that the corrosion current densities (Icorr)
increase and the corrosion potential values (Ecorr) are moved toward negatives values.
The inhibition efficiency E% can be calculated by the following equation:

I'corr
E%  (1  )x100 (1)
I corr

Where Icorr and I’corr are, respectively, the uninhibited and inhibited corrosion current densities
determined by extrapolation of the cathodic Tafel lines to corrosion potential (Ecorr).
Int. J. Electrochem. Sci., Vol. 6, 2011 1459

Table 2. Corrosion parameters obtained from potentiostatic polarization at various temperatures


studied for tinplate in 0.5M HCl containing various concentrations of Artemisia oil.

Concentration T Ecorr Icorr βc βa E


(g/l) (K) (mV/SCE) (mV/dec.) (mV/dec.) (mV/dec.) (%)
Blank 298 -560 102 -127 26 -
308 -563 124 -96 15 -
318 -567 127 -91 16 -
328 -566 172 -90 19 -
338 -580 221 -70 15 -
0.05 298 -599 42 -82 22 59
308 -610 66 -65 15 47
318 -609 70 -42 14 45
328 -612 98 -30 13 43
338 -610 132 -23 10 40
0.1 298 -551 25 -113 24 75
308 -543 44 -106 17 65
318 -552 49 -99 14 61
328 -563 86 -46 16 50
338 -590 105 -103 15 52
0.5 298 -539 19 -119 26 81
308 -546 32 -94 20 74
318 -545 41 -56 18 68
328 -547 57 -35 15 67
338 -559 90 -34 14 59
1 298 -570 10 -99 19 90
308 -565 24 -48 16 81
318 -559 37 -32 14 71
328 -557 51 -26 13 70
338 -571 82 -18 10 63
4 298 -493 8 -106 23 92
308 -490 20 -91 22 84
318 -492 31 -101 19 76
328 -510 46 -78 22 73
338 -531 77 -73 20 65
6 298 -565 5 -95 19 95
308 -543 15 -67 14 88
318 -536 27 -60 21 79
328 -545 41 -30 16 76
338 -542 62 -75 19 72

It’s known that the inhibition efficiency depends on the molecular structure of the
inhibitor and the adsorption of the molecule on the metal surface occurs through the hetero-
atoms such as oxygen (O). If we go back to Fig.1, we find that all molecules of AO components
contain one hetero-atom of oxygen with the exception of Camphene component. Moreover, the
presence of oxygen atom may increase the polarity and the absorbability of the oil on the
surface of tinplate. For this reason, the molecule of camphene doesn’t participate in the
phenomenon of adsorption. Thus, the inhibiting effect of AO may be attributed essentially to
camphor and α-thuyone which are the major components of the oil.
The variation of the logarithm of Icorr of tinplate in 0.5M HCl at various concentrations of AO
as function of temperature reciprocate was illustrated in Fig. 3.
Int. J. Electrochem. Sci., Vol. 6, 2011 1460

Figure 3 shows that the corrosion reaction can be regarded as an Arrhenius-type process
(Equation 2). The activation parameters for the studied system (Ea, H* and S*) were estimated from
the Arrhenius equation and transition state equation (Equation 2 and 3):

E
I cor  A exp(  a ) (2)
R.T

I  RT .exp( S *).exp(  H *) (3)


Nh R RT

where A is Arrhenius factor, Ea is the apparent activation corrosion energy, N is the


Avogadro’s number, h is the Plank’s constant, R is the perfect gas constant, H* and S* are the
enthalpy and the entropy changes of activation corrosion energies for the transition state complex.
Transition state plots for the corrosion rates (Icorr) of the tinplate in the absence and presence of
different concentrations of the inhibitor are given in Fig. 4 and the corresponding activation parameters
Ea, H* and S* for the corrosion process were estimated and listed in Table 3. The change in the
activation free energy (G*) of the corrosion process can be calculated at each experimental
temperature by applying the much known equation:

4
Ln(Icor)

Blank
3 0.05 g/l
0.1 g/l
0.5 g/l
2
1g/l
4 g/l
1 6g/l

0
2,9 3 3,1 3,2 3,3 3,4
1000/T (K-1)

Figure 3. Arrhenius plots of ln(Icorr) versus 1/T at various concentrations of AO.


6

4
Ln(Icor)

Blank
3 0.05 g/l
0.1 g/l
0.5 g/l
2
1g/l
4 g/l
1 6g/l

0
2,9 3 3,1 3,2 3,3 3,4
1000/T (K-1 )

Figure 4. Arrhenius plots of ln(Icorr/T) versus 1/T at various concentrations of AO.


Int. J. Electrochem. Sci., Vol. 6, 2011 1461

G* = H* − TS* (4)

The obtained G*values was also listed in Table 3.

Table 3. The value of activation parameters Ea, H*, S* and G*, and for tinplate in 0.5M HCl in
the absence and presence of different concentrations of AO.

Concentration Ea H* S* G*


(g/l) (kJ.mol-1) (kJ.mol-1) (J.mol-1.K-1) (kJ.mol-1.K-1) at
298 K
Blank 15.57 12.93 -163.23 61,57
0.05 22.47 19.84 -146.78 63,58
0.1 29.68 27.05 -126.78 64,83
0.5 30.86 28.23 -125.29 65,56
1 41.73 39.10 -93.14 66,85
4 45.07 42.44 -83.86 67,43
6 50.93 48.30 -67.39 68,38

From table 3, it can be concluded that:

 As observed, the trend of Ea for the studied inhibitor is not the same with that obtained from
inhibition efficiency. The higher activation energy for oil as compared to that of free acid may be
explained according to Riggs and Hurd [30], as they stated that at higher levels of surface coverage
the corrosion process may proceed on the adsorbed layer of inhibitor and not on the metal surface
leading to a decrease in the apparent activation energy and in some cases become less than that
obtained in the absence of inhibitor.
 The positive signs for both Ea and H* reflecting the endothermic nature of the corrosion process.
It is obviously that the activation energy strongly increases in the presence of the inhibitor. Some
authors [31-33] have attributed this result to the inhibitor species being physically adsorbed on the
metal surface. In this respect, the comparison of the inhibiting action of the investigated
compounds in HCl will be of definite interest.
 The negative values of S* pointed to a greater order produced during the process of activation.
This can be achieved by the formation of an activated complex representing an association or
fixation with consequent loss in the degrees of freedom of the system during the process [34].
 The G* values for the inhibited systems were more positive than that for the uninhibited systems
revealing that in presence of inhibitor addition the activated corrosion complex becomes less stable
as compared to its absence [30].

The fraction of the surface covered by adsorbed molecules of the oil (θ) was determined by the
ratio E%/100. Data were tested graphically by fitting to various isotherms. The plot of (C/θ) against the
inhibitor concentration (C) yields a straight line with correlation coefficient equal unity (Fig. 5).
Int. J. Electrochem. Sci., Vol. 6, 2011 1462

The strong correlation shows that Artemisia essential oil tested was adsorbed on the tinplate
surface electrode according to the Langmuir isotherm [33]:

C 1
 C (5)
θ K

1 ΔG ads
K exp(  ) (6)
55.5 RT

10
9
8
7
6
C/ q

5 298 K
4 308 K
3 318 K
2 328 K

1 338 K

0
0 1 2 3 4 5 6 7
C (g/l)

Figure 5. Curves fitting of the corrosion data for tinplate in 0.5M HCl in the presence of Artemisia oil
according to Langmuir thermodynamic kinetic model at various temperatures.

In Figure 6 shows the evolution of Ln (K) versus the inverse of temperature.

3,2

2,8
Ln(K)

2,6

2,4

2,2

2
2,9 3 3,1 3,2 3,3 3,4
-1
1000/T (K )

Figure 6. Variation of ln(K) versus 1/T.


Int. J. Electrochem. Sci., Vol. 6, 2011 1463

Hads found by Van't Hoff equation can also be calculated by the equation of Gibbs-Helmotz:

G H
ads  ads  cte
T T

The Figure 7 represents the variation of Gads/T versus the inverse of temperature.

-0,050
 Gads /T (KJ/mol.K)

-0,052

-0,054

-0,056

-0,058

-0,060
2,90 3,00 3,10 3,20 3,30 3,40
-1
1000/T (K )

Figure 7. Relation between Gads/T et 1/T

The enthalpy of AO oil adsorption on the tinplate surface can be deduced from the following
equation:

Gads = Hads − TSads (7)

All the results are summarized in Table 4.

Table 4. The value of Activation parameters Hads Sads and Gads for tinplate in 0.5M HCl in the
absence and presence of different temperature.

T (K) K Gads Hads Sads


(kJ.mol-1) (kJ.mol-1) (J.mol-1)
298 22.37 -17.64 -14.8 9.70
308 15.31 -17.26 8.16
318 14.31 -17.64 9.10
328 14.39 -18.22 10.56
338 9.56 -17.62 8.50
Int. J. Electrochem. Sci., Vol. 6, 2011 1464

The negative values of Gads and the higher values of K reveal the spontaneity of adsorption
process and are characteristic of strong interaction and stability of the adsorbed layer with the tinplate
surface. The obtained values of the adsorption free energy were >- 36 to - 44 kJ mol-1 are indicative of
chemisorption 35 [35]. The adsorption of an organic adsorbate at the metal/solution interface is
considered a ‘‘substitutional adsorption’’ phenomenon [36]. Therefore, the positive values of Sads and
Hads related to ‘‘substitutional adsorption’’ can be attributed to the increase in the solvent entropy and
to a more positive water desorption enthalpy.

3.3. Electrochemical impedance spectroscopy measurements

Fig. 8-9 shows EIS measurements obtained in 0.5 M HCl in the absence and presence of 0.5g/l
of Artemisia essential oil at various immersion times.

150
-Zi(ohm. cm )
2

100
0.5h
2h
50 4h
24h
0
0 100 200 300 400
2
Zr(ohm. cm )

Figure 8. Nyquist plot for tinplate in 0.5 M HCl of different time of immersion at 298K
1000

800
-Zi(ohm.cm )

600
2

400
0.5h
2h
200
4h
24h
0

0 1000 2000 3000 4000


2
Zr(ohm.cm )

Figure 9. Nyquist plot for tinplate in 0.5 M HCl in the presence of different time of immersion of
0.5 g/l AO at 298 K.

Inspection of Fig. 8 and 9 clearly shows that AO as a corrosion inhibitor for tinplate in 0.5 M
HCl. Similar figures have been described in the literature for the acidic corrosion of iron and steel in
the presence of inhibitor molecules [37-38].
Int. J. Electrochem. Sci., Vol. 6, 2011 1465

It can be observed that the size of the capacitive arc in the presence of AO is larger than that in
the absence of this inhibitor. This fact is attributed to the inhibition effects of the AO on tinplate
corrosion.
Table 5 shows the charge transfer resistance, double layer capacitance and calculated
efficiencies obtained from Nyquistic diagram for 0.5 g/l of AO at 298 K.

The inhibition efficiency can be calculated by the following formula:

( Rt  R t )
E Rt %  x100
Rt

Here Rt and R0t are the charge transfer resistances in uninhibited and inhibited solutions
respectively.
The values of the polarization resistance were calculated by subtracting the high frequency
intersection from the low frequency intersection [39]. Double layer capacitance values were obtained
at maximum frequency (fm), at which the imaginary component of the Nyquist plot is maximum, and
calculated using the following equation.

1
C  (7)
dl 2. . f .R
m t

With Cdl : Double layer capacitance (µF.cm-2) ; fm : maximum frequency (Hz) and Rt : Charge
transfer resistance (Ω.Cm2)

Table 5. Impedance parameters for the corrosion of tinplate in 0.5 HCl solution in the absence and
presence of AO at 0.5 g/l.

Immersion time Rt Cdl ERt %


(hours) (Ω.Cm2) (µF.cm-2)
Blank 0.5 223.70 89.65 94
0.5g/l 4258.00 0.37
blanc 2 90.93 875.59 95
0.5g/l 2213.00 0.91
blanc 4 47.45 1766.24 93
0.5g/l 840.70 4.24
blanc 24 32.49 3101.94 94
0.5g/l 670.40 59.38

The examination of data of this table may these conclusions:


1. The nhibitory efficiency is practically independent on immersion time.
2. Values of Cdl in inhibited solution are lower than those in uninhibited acid.
Int. J. Electrochem. Sci., Vol. 6, 2011 1466

3.4. SEM measurements

Fig. 10 shows a scanning electron microscope (SEM) photomicrograph of tinplate surface


before and after immersion in 0.5M HCl with and without the addition of the inhibitor at 298 K. The
specimen surface in Fig. 10b appears to be roughened extensively by the corrosive environment and
the porous layer of corrosion product is present. In presence of AO inhibitor, the data (Fig. 10c) gave
the formation of thick films on tinplate surface. This may be intercepted by the adsorption of this
inhibitor on the electrode surface. It is thought that the molecule of Aremisia oil depresses the
corrosion by the formation of a protective film on the electrode surface.

(A) (B) (C)

Figure 10. SEM (x1000) of tinplate (a) before immersion (b) after 24 hours of immersion in 0.5M HCl
(C) after 24 hours of immersion in 0.5M HCl + 0.5 g/l of AO.

4. CONCLUSION
In this work, we have studied the behaviour of corrosion of tinplate in hydrochloride media
taking account the effect of the addition of natural Artemisia essential oil at different concentrations
and the variation of medium temperature. From the obtained results, we can have the following
conclusions:

 Increasing of temperature stimulates the active dissolution of tinplate layer in HCl medium.
 Addition of AO reduces the corrosion rate of tinplate.
 The inhibitive action of AO increases with concentration.
 The best resistance to the corrosion gotten is 81% at temperature 298 K at 0.5 g/l AO
concentration.
 Adsorption of inhibitor tested fellows Langmuir adsorption isotherm.

References

1. Ait Addi. E., Bazzi. L., El Hilali. M., Salghi. R., Hammouti. B., Mihit. M. Appl. Surface Sci. 253
(2006) 555-560.
2. Ait Addi. E., Bazzi. L., El Hilali. M., Zine. E., Salghi. R., El Issami. S. Can. J. Chem. 81 (2003)
297.
3. Foad El-Sherbini. E.E. J. of Electroanalytical Chem. 584 (2005) 167.
Int. J. Electrochem. Sci., Vol. 6, 2011 1467

4. Aride. J., Ben-Bachir. A., Elkacemi. K., Etman. M., Kertit. S., Srhiri. A. J. Chim. Phys. Physico-
Chimie biologique. 88 (1991) 631.
5. Bastidas, J.M., Cabanes, J.M., Catala, R.J. Coating Technol. 67 (1997) 69.
6. Abd El Rehim. S.S., Hassan. H.H., Mohamed. N.F. Corros. Sci. 46 (2004) 1071.
7. Refaey. S.A.M. Electrochim. Acta. 41 (1996) 2545.
8. Foad El-Sherbini. E.E.S., Abd-El-Wahab. M., Amin. M.A., Deyab. M.A. Corros. Sci. 48 (2006)
1885.
9. Drogowska. M., Brossard. L., Ménard. H. J. of appl. Electrochem. 19 (1989) 231.
10. Belkhaouda, M., Bazzi, L., Balhachemi, A., Salghi, R., Hammouti, B., kertit, S. Applied Surface
Sci. 252 (2006) 7921.
11. Bammou, L. Thèse de doctorat, Ecole Nationale des Sciences appliquées, Agadir, (2010).
12. El-Etre. A.Y., Abdallah. M., El-Tantawy. Z.E. Extract. Corros. Sci. 47 (2005) 385-395.
13. Bothi. R.P., Sethuraman. M.G. Mater. Lett. 62 (2008) 2977-2979.
14. Bouyanzer. A., Majidi. L., Hammouti. B. Phys. Chem. News. 37 (2007) 70-74.
15. Chetouani. A., Hammouti. B., Benkaddour. M. Pigm. Resin Technol. 33 (2004) 26-31.
16. Okafor. P.C., Ikpi. M.E., Uwah. I.E., Ebenso. E.E., Ekpe. U.J., Umoren. S.A. Corros. Sci. 50
(2008) 2310-2317.
17. Noor. E.A. J. Eng. Appl. Sci. 3 (2008) 23-30.
18. Benabdellah. M., Benkaddour. M., Hammouti. B., Bendahhou. M., Aouniti. A. Appl. Surf. Sci. 252
(2006) 6212-6217.
19. Bouyanzer. A., Hammouti. B., Majidi. L. Mater. Lett. 60 (2006) 2840-2843.
20. Oguzie. E.E. Portugaliae Electrochim. Acta. 26 (2008) 303-314.
21. Oguzie. E.E. Azadiracta indica. Corros. Sci. 50 (2008) 2993-2998.
22. Oguzie. E.E., Iyeh. K.L., Onuchukwu. A.I. Bull. Electrochem. 22 (2006) 63-68.
23. Gong. M., Mao. F-Y., Wu. J-P., Zeng. X-G. Corros. Prot. 27 (2006) 576-578.
24. Bouyanzer. A., Majidi. L., Hammouti. B. Bull. Electrochem. 22 (2006) 321-324.
25. Bammou. L. Chebli. B., Salghi. R. , Bazzi. L., Hammouti. B., Mihit. M. and El Idrissi. H. Green
Chem. Lett. and Rev. In press 2010
26. Anonymous. French Pharmacopoeia. Maisonneuve SA: Moulins-les-Metz. France. 1983.
27. Stenhagen. E., Abrahamsson. S., McLafferty. F.W. Registry of Mass Spectral Data. John Wiley:
New York. NY. 1974.
28. Adams. R.P. Allured: Carol Stream. IL. 1995.
29. Shreir. L.L., Jarman. R.A., Burstein. G.T. Corrosion: Metal/Environment Reactions. Edition:
Butterworth-Heinemann. 1 (2000) 4:159.
30. Riggs. J.L.O., Hurd. T.J. Corros. (Nace) 23 (1967) 252-258.
31. Almeida. C.M.V.B., Raboczkay. T., Griannetti. B.F. J. Appl. Electrochem. 29 (1999) 123-128.
32. Bentiss. F., Lebrini. M., Lagrenee. M. Corros. Sci. 47 (2005) 2915.
33. Mora. N., Cano. E., Polo. J.L., Puente. J.M., Bastidas. J.M. Corros. Sci. 46 (2004) 563-578.
34. Bouklah. M., Hammouti. B., Lagrenee. M., Bentiss. F. Corros. Sci. 48 (2006) 2831-2842.
35. Szlarska-Smialowska, Z. Corros. Sci. 18 (1978) 557.
36. Branzoi.V., Branzoi. F., M. Baibarac, Mater. Chem. Phys. 65 (2000) 288.
37. Yadav A.P., Nishikata A., Tsurn T., Corros. Sci. 46 (2004) 104.
38. Bentiss F., Traisnel M., Gengembre L., Lagrenee M., Appl. Surf. Sci. 152 (1999) 237.
39. Elachouri M., Infante M.R, Izquerdo F., Kertit S., Gouttaya H.M., Nciri B., Corros. Sci. 43 (2001)
19.

© 2011 by ESG (www.electrochemsci.org)

You might also like